Carbene Complexes. Catalytic Metathesis of Azobenzene De

be difficult to make otherwise. Rather than involving metal nitrenes, the reaction proceeds via aza-ylides that evolve into diaziri- dines; a meta-sta...
8 downloads 26 Views 500KB Size
Subscriber access provided by University of Florida | Smathers Libraries

Article

Structure and Reactivity of Half-Sandwich Rh(+3) and Ir(+3) Carbene Complexes. Catalytic Metathesis of Azobenzene Derivatives Daniel J. Tindall, Christophe Werlé, Richard Goddard, Petra Philipps, Christophe Farès, and Alois Fürstner J. Am. Chem. Soc., Just Accepted Manuscript • DOI: 10.1021/jacs.7b12673 • Publication Date (Web): 14 Jan 2018 Downloaded from http://pubs.acs.org on January 14, 2018

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

Journal of the American Chemical Society is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 10 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of the American Chemical Society

Structure and Reactivity of Half-Sandwich Rh(+3) and Ir(+3) Carbene Complexes. Catalytic Metathesis of Azobenzene Derivatives Daniel J. Tindall, Christophe Werlé, Richard Goddard, Petra Philipps, Christophe Farès, and Alois Fürstner* Max-Planck-Institut für Kohlenforschung, D-45470 Mülheim/Ruhr, Germany ABSTRACT: Traditional rhodium carbene chemistry relies on the controlled decomposition of diazo derivatives with [Rh2(OAc)4] or related dinuclear Rh(+2) complexes, whereas the use of other rhodium sources is much less developed. It is now shown that halfsandwich carbene species derived from [Cp*MX2]2 (M = Rh, Ir; X = Cl, Br, I, Cp* = pentamethylcyclopentadienyl) also exhibit favorable application profiles. Interestingly, the anionic ligand X proved to be a critical determinant of reactivity in the case of cyclopropanation, epoxide formation and the previously unknown catalytic metathesis of azobenzene derivatives, whereas the nature of X does not play any significant role in -OH insertion reactions. This perplexing disparity can be explained on the basis of spectral and crystallographic data of a representative set of carbene complexes of this type, which could be isolated despite their pronounced electrophilicity. Specifically, the donor/acceptor carbene 10a derived from ArC(=N2)COOMe and [Cp*RhCl2]2 undergoes spontaneous 1,2-migratory insertion of the emerging carbene unit into the Rh−Cl bond with formation of the C-metalated rhodium enolate 11. In contrast, the analogous complexes 10b,c derived from [Cp*RhX2]2 (X = Br, I) as well as the iridium species 13 and 14 derived from [Cp*IrCl2]2 are sufficiently stable and allow true carbene reactivity to be harnessed. These complexes are competent intermediates for the catalytic metathesis of azobenzene derivatives, which provides access to α-imino esters that would be difficult to make otherwise. Rather than involving metal nitrenes, the reaction proceeds via aza-ylides that evolve into diaziridines; a meta-stable compound of this type has been fully characterized.

INTRODUCTION Dirhodium carbenes A formed in situ upon reaction of a diazo derivative with a (chiral) dinuclear Rh(II) tetracarboxylate complex are exceptionally versatile synthetic intermediates that power a host of transformations, including (asymmetric) cyclopropanations, C−H insertions, cycloaddition reactions and a myriad of ylide chemistry (Scheme 1).1-7 Although the original report on the use of [Rh2(OAc)4] had indicated that a salt as simple as RhCl3⋅3H2O is also catalytically competent,8a mononuclear Rh(+3) carbene complexes in general and halfsandwich carbenes derived from [Cp*RhX2] precursors in particular have not been studied nearly as thoroughly.8a,9 Only recently have intermediates of this type been implicated in directed C−H activation reactions, in which the precatalyst is first ionized e. g. with a soluble Ag(I) salt; the resulting electrophilic metal fragment [Cp*Rh]2+ then (cyclo)metalates the (arene) substrate prior to reaction with a diazo derivative to give a putative Rh(+3) carbene of type C which undergoes migratory insertion to give an ortho-substituted product.10 This sequence has become quite popular: although most applications use transient carbenes C with at least one acceptor substituents (R1 and/or R2 = electron withdrawing group),11 the example from our laboratory shown in Scheme 1 proves that even a donor/donor carbene is sufficiently electrophilic to give the meta-stable metallacycle 3 (Ar = MeOC6H4-) comprising a rare tert-alkyl rhodium unit, which could be characterized by crystallographic means (Figure 1).12 The fact that complex 4 as the neutral analogue of the purported intermediate 2 could also

be isolated in crystalline form (for its structure in the solid state, see Figure S1) lends credence to the proposed pathway.12 In contrast to the popularity of this type of directed C-H activation via cyclometalated rhodium carbenes C, proper pianostool Cp*Rh(+3) carbene complexes of type B have been studied much less extensively.13 This situation, however, might be unjustified: During recent studies on structure and bonding in late-transition metal carbene complexes,14-16 we found such species to be competent intermediates in catalytic cyclopropanation, alkoxyfuran formation, as well as RO−H and R3Si−H insertion reactions.17 Moreover, crystal structures of two reactive complexes of type B have been obtained.17 We now follow up on these initial results and disclose a more comprehensive investigation into half-sandwich Cp*M carbene complexes of Rh(+3) and Ir(+3). Most notably, a striking correlation between the chosen counterion X and the stability and reactivity of the corresponding carbene was noticed that can be explained on the basis of the acquired structural data. Moreover, we now report that the new pianostool carbene complexes enable catalytic metathesis reactions of azo compounds which seem to be unprecedented in the literature. Scheme 1. Prototype Rhodium Carbene Complexes (top); Example of a “Donor/Donor” Carbene Participating in Directed CH-Activation/Carbene Insertiona

ACS Paragon Plus Environment

Journal of the American Chemical Society R

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

O O R

O

Rh

O O R

O Rh O

X

R O

X

R2

Rh

R1

A

B

a)

R1

C

Ar Ar

SbF 6 N N

Ph

R2 Z

R1

Rh

HN

Rh

R2

X

Rh

SbF6

N R1

b)

Ph

R 2HN 3 R1 = Ph, R2 = H 3' R 1 = H, R 2 = Ph

2 NH2

NH c)

N2

1 Rh

cis-disposition of the ester and the aryl substituent (see Figure S2) implies intervention of the ylide conformer E which is favored over D on steric grounds as it avoids the substituent R from clashing into the bulky ligand sphere about Rh comprising the Cp* and two chloride ligands. Epoxide formation proceeded without incident even if an electron-withdrawing substituent on the arene ring renders the aldehyde oxygen atom less nucleophilic and hence carbonyl ylide formation less favorable (entry 6).23 Overall, [Cp*RhI2]2 compares favorably with Rh2(OAc)4 previously used in the literature (compare entries 2/3 and 7/8), whereas [Cp*RhCl2]2 is inadequate.19 As previously communicated, [Cp*RhI2]2 also performed well in a Si−H insertion reaction;17 once again, use of the iodidecontaining catalyst is necessary, whereas its chloride-based analogue basically fails to react.

Cl N

4

Page 2 of 10

NH2

Ph

X

5a , X = OMe 5b, X = NMe2

X

Table 1. Catalyst Screening: Cyclopropanation X

Reagents and Conditions: a) (i) [Cp*Rh(MeCN)3][SbF6]2, MeCN, reflux, see ref. 12; b) 5a, CH2Cl2; c) [Cp*RhCl2]2, NaOAc, toluene, reflux; Ar = 4-methoxyphenyl

RO

O

a

O

MeO OR

N2

X

Entry

Substrate

Catalyst (1 mol%) 1 [Cp*RhCl2]2 6a 2 [Cp*RhI2]2 6a 3 [Cp*IrCl2]2 6a 4 [Cp*IrI2]2 6c a Yield of isolated product

OMe

7

6a, X = OMe, R = Me b, X = H, R = Me c, X = COOMe, R = Et

Solvent pentane pentane CH2Cl2 CH2Cl2

T (°C) 22 22 40 22

Yield (%)a ≤ 46 79 63 58

Table 2. Catalyst Screening: Darzens-type Epoxide Formationa O

X

X

O

O

X

R

Figure 1. Structure of the cyclometalated product 3 formed by directed C−H activation followed by carbene insertion; the disordered SbF6− counterion and solute CH2Cl2 are not shown for clarity

6 R

Nr

RESULTS AND DISCUSSION Strong Effect of the Anionic Ligand on Reactivity. The reaction of the diazo derivative 6a with excess 4methoxystyrene was chosen for an initial catalyst screening (Table 1). [Cp*RhCl2]2 (1 mol%) afforded the expected cyclopropane 7, but the yield remained low (≤ 46%) despite attempted optimization. Surprisingly, the result was improved when the chloride ligands on rhodium where replaced by iodide. Similarly, it was found that the formal substitution of Rh by Ir as the central metal furnished 7 in appreciable yield;18 although the chosen anionic ligands still exert some influence, the effect seems to be less pronounced in the Ir series. The same trend was observed in reactions with aromatic aldehydes, which are known to give epoxides when reacted with certain metal carbenes generated in situ.19-22 Because the reaction is thought to commence with attack of the C=O heteronucleophile onto the carbene to give a carbonyl ylide intermediate, we conjectured that Rh(+3) half-sandwich carbenes might be particularly well suited because of their high oxidation state (Table 2). To our surprise, however, [Cp*RhCl2]2 led to hardly any conversion (entry 1), whereas [Cp*RhI2]2 afforded 8 in good yield as a single diastereomer (entry 2). The

Substrate

[Rh]

O

[Rh]

D

O

OMe

OMe

E

R

O

COOMe R

8

R

Catalyst Yield (1 mol%) (%)c 1 Ph [Cp*RhCl2]2 7 (GC) 6b 2 Ph [Cp*RhI2]2 61 6b 3 Ph Rh2(OAc)4 b 6619b 6b 4 4-MeOC6H4[Cp*RhI2]2 96 6b 5 4-MeOC6H4[Cp*RhI2]2 83 6a 6 4-(MeOOC)C6H4[Cp*RhI2]2 85 6b 7 PhCH=CH[Cp*RhI2]2 88 6b 8 PhCH=CHRh2(OAc)4 b 5019b 6b 10 6b EtCH=CH[Cp*RhI2]2 98 a Unless stated otherwise, all reactions were carried out in CH2Cl2 at ambient temperature; b in refluxing CH2Cl2; c yield of isolated product, unless stated otherwise

The strong correlation between efficiency and the nature of the halide ligand X manifest in these examples stands in striking contrast to the results for the catalyzed insertion of 6 into the O−H bond of methanol (Table 3): all tested Cp*MX2 precatalysts − independent of their anionic ligand − afforded product 9 in good to excellent yield; even an Ir(I) complex proved effective. Table 3. Catalyst Screening: -OH Insertion Reactiona

ACS Paragon Plus Environment

Page 3 of 10

Journal of the American Chemical Society X

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

MeOH (10 equiv.)

O OR

6

N2

X

O OR

CH2 Cl2, RT

9

OMe

Yield (%)b Entry X Catalyst (1 mol%) c 1 [Cp*RhCl ] 83 −OMe 2 2 2 88 [Cp*RhI2]2 −OMe 3 [Cp*IrCl2]2 74 −OMe 4 [Cp*IrI2]2 76 −OMe 5 [Cp*IrI2]2 90 −COOMe 6 [(cod)IrCl]2 98c −OMe a All reactions were carried out in CH2Cl2 or pentane at ambient temperature; b yield of isolated product, unless stated otherwise; c NMR yield

placement of the donor/acceptor entity of transient 10a by an inherently less electrophilic donor/donor carbene15a,30 would also prevent such a spontaneous insertion step from occurring. In fact, reaction of [Cp*RhCl2]2 with 5b furnished complex 12, in which both chlorides remain bound to the Rh(+3) center and the carbene unit is fully intact (Scheme 2 and Figure 2). As one might expect, the Rh1−C1 distance (2.014(3) Å) of 12 is somewhat longer than that of 10c (1.970(3) Å).17 Scheme 2. Preparation of Half-Sandwich Rh(+3) Carbenesa 5b

6a [Cp*RhX2]2

X

a)

These strikingly different trends could not be reconciled at the outset of the project. Therefore it was deemed necessary to characterize the purported half-sandwich carbene intermediates of type B in more detail. Structures of Cp*M Carbene Complexes (M = Rh, Ir) in the Solid State. Despite the tremendous preparative importance, carbene complexes A derived from dirhodium tetracarboxylate precatalysts had defied all attempts at experimental characterization for decades; only recently have spectral and crystallographic data of a representative set of reactive intermediates of this type been obtained.15,17,24 In concord with computational results,25 the gathered information suggests that the second rhodium atom exerts a stabilizing function as an integral part of the three center/four electron Rh(+2)−Rh(+2)−CR2 core.26 In recognition of the critical role of metal-metal bonding, it was expected that mono-nuclear complexes of type B might be even more fragile because such assistance by a neighboring metal center is missing and the oxidation state of rhodium is also higher (+3 in B versus +2 in A). In fact, it required considerable experimentation until we managed to isolate the donor/acceptor carbene complexes 10 derived from the diazo ester derivative 6a and [Cp*RhX2]2 (X = Cl, Br, I) (Scheme 2). As previously reported, compounds 10b (X = Br) and 10c (X = I) represent the first true halfsandwich donor/acceptor carbene complexes of Rh(+3) to be fully characterized.17 In contrast, the chloride analogue 10a (X = Cl) escaped all attempts at isolation: although there is no doubt that this species is transiently formed, the nascent carbene unit inserts into one of the Rh−Cl bonds even at −50°C.27,28 The structure of the resulting product 11 in the solid state shows a functionalized C-metalated rhodium enolate;29 alternatively, this species can be viewed as a rhodium carbenoid because it comprises a C-atom carrying a metal atom and chloride as a potential leaving group.17 The metalated center resonates at δC = 70.6 ppm (CD2Cl2, −50°C), whereas the true carbene complexes 10b,c have the expected lowfield NMR signatures (δC = 314.2 ppm (X = Br), 316.4 ppm (X = I)). Interestingly, these resonances are significantly downfield from those of all dirhodium carbene complexes of type A characterized so far (ca. 235-270 ppm).17,24 The spontaneous formation of 11 at cryogenic conditions suggests that the chloride ligands render the carbene center of transient 10a so electrophilic that spontaneous 1,2-migratory insertion will ensue; in contrast, the somewhat less electronegative bromide or iodide ligands are incapable of triggering this process. Under this premise it seemed likely that formal re-

b)

OMe

Rh

X MeO

X = Cl

[10a , X = Cl]

O

10b, X = Br 10c , X = I

X = Cl

Cl

NMe2

Rh

Cl

OMe Rh Cl MeOOC Cl

12

Me 2N

11

Reagents and conditions: (a) 6a, CH2Cl2/FC6H5, 0°C → −20°C, see ref. 17; (b) 5b, CH2Cl2/FC6H5, 0°C → −20°C a

Figure 2. Structure of the donor/donor Cp*Rh(+3) carbene complex 12 in the solid state

Scheme 3. Preparation of Iridium-Carbene Complexesa O X

Ir

OMe

X EtO

6c b)

6a a)

Cl

OMe

Ir

Cl MeO

14a , X = Cl b, X = I

O

O

Ir

6a [(cod)IrCl] 2

[Cp*IrX2 ]2

Cl COOMe

Ir

13

Cl COOMe

c)

15 OMe

OMe

Reagents and Conditions: a) 6a, CH2Cl2/FC6H5, 0°C → −20°C; (b) 6c, CH2Cl2/FC6H5, 0°C → −20°C; c) 6a, CH2Cl2/FC6H5, 0°C → −20°C; cod = 1,5-cyclooctadiene a

ACS Paragon Plus Environment

Journal of the American Chemical Society 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 3. Structure of the Ir(+3) half-sandwich carbene complex 13 in the solid state; disordered fluorobenzene in the unit cell is not shown for clarity

Page 4 of 10

NMR spectra of 15 are also broad due to hindered rotation about the C1−C4 bond. The partial double bond character manifest in this spectral feature corresponds well to the crystallographic data, which show a short C1−C4 distance and an almost perfect alignment of the arene with the carbene “empty” orbital (Figure 4). As in 13, it is largely the arene substituent that imparts (meta)stability onto this donor/acceptor carbene via delocalization of charge density within the aromatic π-cloud.15,17 In line with the lower oxidation state of and hence higher electron density at the central metal, the Ir1−C1 carbene bond (1.912(3) Å) of 15 is shorter than that of the Ir(+3) species 13 (1.960(2) Å).34 Yet, the dynamic behavior observed by NMR proves that even in 15 the carbene unit has little double bond character; the bond order of its Ir(+3) sibling 13 must therefore be even lower. Not unexpectedly, the preparation of complexes 14a,b was even more challenging as they carry an electron-withdrawing ester function (rather than a MeO- substituent) on the aryl ring (Figure 5).35 The emerging “acceptor/acceptor” character is manifest in the structural data of 14a in the solid state: as a deactivated arene is obviously less capable of compensating the electron deficiency of the adjacent carbene, the C1−C5 bond (1.460(4) Å) is markedly longer than that in the donor/acceptor carbene 13 (1.429(3) Å). As a consequence, back-donation from [Cp*Ir] is upregulated, which results in a significant contraction of the carbene Ir1−C1 bond to 1.914(3) Å in 14a (versus 1.960(2) Å in 13).

Figure 4. Structure of the Ir(+1) donor/acceptor carbene complex 15

Formal replacement of Rh(+3) by Ir(+3) as the central metal also suppressed the ligand coupling (Scheme 3). Complex 13 is the iridium analogue of the transient rhodium carbene 10a that had defied characterization (Figure 3).31 Interestingly, the carbene center of 13 resonates at δC = 275.9 ppm (CD2Cl2) and is hence much less deshielded than those of the rhodium species 10b (314.2 ppm) and 10c (316.4 ppm). Moreover, the Ir1−C1 carbene bond (1.960(2) Å) is shorter than the Rh1−C1 bonds in 10c (1.970(3) Å),17 which speaks for stronger backbonding from Ir(+3) into the “empty” carbene orbital. The arene ring in 13 is aligned for maximal orbital overlap between the π-system and the carbene; as a consequence, the C1−C4 bond is markedly contracted to 1.429(3) Å. These structural attributes showcase that the donor substituent largely accounts for the stabilization of highly electrophilic species of this type. In contrast, the electron-withdrawing ester group is almost orthogonal to the carbene as to prevent any further destabilization of the already highly electron deficient site; the decoupling is also manifest in the rather long C1−C2 distance (1.510(3) Å). We also managed to isolate the analogous Ir(+1) carbene 15; the interest in this species was nurtured by the fact that [(cod)IrCl]2 had proven competent in –OH insertion reactions (Table 3, entry 6).32,33 The 13C NMR spectrum of 15 shows four pairs of differently broadened lines for the cod ligand. This observation suggests twofold symmetry, whereby the broadening comes from the dynamic averaging about the symmetry axis. The resonances of the aromatic signals in the

Figure 5. Structure of the half-sandwich “acceptor/acceptor” carbene 14a in the solid state; CH2Cl2 in the unit cell is not shown for clarity

These insights into the subtle interplay between the central metal and the ligand framework allowed the reactivity trends to be rationalized that had surfaced during the initial screening. The striking correlation between the halide ligand X in the precatalyst and the efficiency of the cyclopropanation of olefins (Table 1) and the Darzens-type reactions of aromatic aldehydes (Table 2) likely has its origins in the distinct migratory aptitudes of X (Cl >> Br ≈ I) and hence the largely different stability of the carbene complexes of type B: the chloridecontaining Cp*Rh(+3) carbene 10a is only transiently formed: intramolecular carbene insertion into the Rh−Cl bond is evidently faster than the intermolecular attack of the C=O group of an aldehyde partner, which prevents carbonyl ylide E from forming and hence brings epoxidation to a halt. The resulting “carbenoid” 11 is also unlikely to cyclopropanate styrene,

ACS Paragon Plus Environment

Page 5 of 10 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of the American Chemical Society

certainly not under mild conditions.28 For 10b,c (M = Rh, X = I, Br) as well as for all Cp*Ir(+3) complexes, in contrast, this migratory process has not been observed; such species are therefore competent partners in the cited transformations. The fact that the reaction with methanol seems largely unaffected by such changes in the precatalyst (Table 3) is understood if one considers that even a carbenoid such as 11 will react with an alcohol substrate.17 Catalytic Azo-Metathesis. The ability of certain Fischer carbene complexes to cleave the −N=N− bond of an azobenzene derivative has previously been demonstrated;36-38 to the best of our knowledge, however, this transformation has never been translated into a catalytic process that would open a relevant new entry into α-imino esters. Compounds of this type find many different applications, which have recently been covered in a comprehensive review;39 therefore it may suffice to mention that asymmetric reduction of α-imino esters provides rapid access to amino acid building blocks, whereas many organometallic reagents attack their imine entity at nitrogen (rather than at carbon) and, in doing so, lead to valuable N-substituted amino acid derivatives.39,40 Although condensation of an α-oxo-ester with the appropriate amine partner seems to be the obvious way for making αimino esters such as 17, numerous alternative methods have been developed in the past.39 Especially when it comes to reacting poorly nucleophilic amines and/or sensitive α-oxo esters, recourse to a phosphine–mediated Staudinger ligation of an azide precursor with the carbonyl is often the method of choice, despite the poor atom economy of this process.41 More recently, α-imino esters were prepared from donor/acceptor dirhodium carbenes upon reaction with aryl azides.42,43 Compared to azides, azoarenes have many obvious advantages as possible source of an [Ar−N=] fragment, because they are stable, safe, benign and readily available in functionalized form; for their role as classical dyes, a sizeable number of such compounds is available even in bulk quantities. Therefore it seemed warranted to investigate whether catalytic metathesis of azobenzene derivatives opens a potentially orthogonal gateway to α-imino esters that are difficult to make otherwise. To this end, however, −N=N− bond cleavage must outperform directed ortho-metalation of the azobenzene substrate.44 This is a non-trivial task since the directing effect of the −N=N− bridge onto rhodium complexes is well established and has already served heterocycle synthesis in the past.45 Gratifyingly though, the reaction of diazo derivative 6b with azobenzene 16 in the presence of catalytic [Cp*RhI2]2 in toluene furnished traces of the desired α-imino ester 17 (Table 4, entry 3; for the product structure, see Figure S3); although a better yield was obtained at reflux temperature, the outcome proved erratic upon attempted scale-up (entry 4). We surmised that light might play a role in triggering isomerization of the largely E-configured azo derivative 16 to the potentially more nucleophilic and sterically less hindered Z-isomer.46,47 This assumption proved correct in that essentially quantitative conversion was reached when the diazo derivative was added over the course of 2 h to the mixture irradiated with light emitted by a commercial blue LED; the 91% yield of pure 17 obtained under these conditions proved well reproducible (entry 8). As no other product derived from the azobenzene substrate was detected in the crude material, both halves of 16 must have ended up in the product. A control experiment using an

unsymmetrical azobenzene derivative confirmed this conclusion (see below, Scheme 4). Table 4. Catalytic Metathesis of Azobenzene

Nr

Catalyst (2 mol%) [Cp*RhCl2]2

solvent

T

LEDa

Yield (%)b 1 toluene RT Cl) when working with half-sandwich rhodium precatalysts of type [Cp*MX2]2; this effect basically vanished for M = Ir(+3) (Table 4). In view of the structural data outlined above, these trends suggest the intervention of true carbene intermediates. In line with this notion, [Rh2(OAc)4] was also found effective (entry 10), although the yield was lower and the functional group tolerance definitely inferior (see below); in contrast, Ir(+1) as well as Cu(+1) proved inadequate (entries 13, 14). Under the optimized conditions, the reaction shows a broad scope and appreciable compatibility with polar and apolar substituents; moreover, it was found to scale well. As can be seen from Table 5, the azobenzene derivative to be metathesized can be electron rich or electron poor, even though the latter usually require higher temperatures. The need for more forcing conditions suggests that the reaction commences with attack of an N-atom of the azo-bridge onto the electrophilic carbene center, which is obviously less favorable for electron deficient substrates. In any case, the new methods allows αimino esters such as 20-22, 25, 28, 33 and 36 to be made, which are difficult to access by conventional carbonyl condensation since the required aniline derivatives are poorly nucleophilic and therefore do not react well with an α-oxo-ester. The substituents can be placed ortho, meta or para to the azo bridge with little change in efficiency; only the 2,6disubstituted derivative 27 did not form, most likely because the −N=N− unit is too shielded for attack onto the rhodium intermediate. Donor/acceptor carbenes of the structural type discussed in the previous sections of this paper gave the best results; electron donating or withdrawing substituents on their aryl rings do not matter much. The more stabilized do-

ACS Paragon Plus Environment

Journal of the American Chemical Society 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

nor/donor carbenes react reluctantly (see product 37): their carbene center is less electrophilic and hence less susceptible to attack by the −N=N− bridge. In contrast, diazo derivatives with only electron withdrawing substituents such as diazomalonate are not decomposed by [Cp*MX2]2 and therefore cannot be used at the current stage of development. Moreover, an attempt at using this reaction for the ring opening of pyridazine, a heteroaromatic “azo”-derivative, met with failure, because this and related substrates form stable adducts with the [Cp*RhCl2] fragment.48

Page 6 of 10

tives reported in the literature had been explained in analogy to canonical olefin metathesis by invoking metal-nitrenes and metallacyclic intermediates.36,37 Although rhodium nitrene complexes could not be excluded as the propagating species at the outset of our investigation, the lack of any scrambling of the azobenzene derivatives 39 and 40 in the control experiment shown in Scheme 4 argued against such a scenario. Scheme 4. Control Experiments Me

COOMe

19 20 21 22

77% (R = OMe) 89% (R = COOC 8H17) b 91% (R = F) 84% (R = CF3)b

R

COOMe N

COOMe CF3

F

MeO

COOMe

Me

N2 Ph COOMe (1.5 equiv.)

N Me COOMe

COOMe

N

N N

N

F

N N

MeO

[Cp*RhI2]2 (1 mol%) LED, toluene, RT

24 99% (R = OMe) 25 86% (R = COOC 8 H 17)

COOMe

38

23 85%

Me N

Ph

COOMe

85% (18:22 , 1:1.05, NMR)

Me

N

Ph

toluene, reflux

N

N

+

[Cp*RhI2]2 (2 mol%)

N

CF3

Me

Ph N

N

N2

2

Table 5. Selection of imines formed by azobenzene metathesisa

not detected

27 0%

26 95%

39

OMe

F

40

O N3 N

N COOMe

N COOMe

28 83% b

Scheme 5. Formation and Evolution of a Diaziridine Intermediate N

COOMe

29 65%

N NOE

30 71%

O

(Z)-16 I

R

R

MeO

N

OMe

Rh

N CD2Cl2, -40°C

I MeO

N COOMe MeO

O

N

N

10c

MeO

41

COOEt

MeO

31 97% (R = H) 32 69% (R = OMe) 33 65% (R = COOC 8 H17) b

10c O 34 90% (R = H) 37 67% e 35 66% (R = OMe) 36 92% (NMR, R = COOC8H 17)b,c,d

2

41

Ph COOMe

-40°C to RT MeO

a

unless stated otherwise, all compounds were made using [Cp*RhI2]2 (1 mol%) in toluene at ambient temperature under irradiation with light emitted by blue LED’s; b at 90°C; c using [Cp*IrI2]2 as the catalyst; d the product is subject to partial hydrolysis during flash chromatography; e addition of the diazo derivative over 5 h

Compounds 28-30 nicely illustrate the favorable chemoselectivity profile of this new transformation. An imine such as 28 bearing an unprotected ketone on its backbone is neither accessible from an α-oxo-ester by ordinary condensation chemistry nor by Staudinger ligation. Furthermore, the intact alkyne in compound 29 shows that catalytic −N=N− cleavage is faster than cyclopropene formation.17 Arguably most notable, however, is the synthesis of product 30: the fact that [Cp*RhI2]2 cleaved the azo bridge while leaving the azide intact sets this catalyst apart from many other metal complexes (including [Rh2(OAc)4] and other dirhodium tetracarboxylate complexes), which readily react with azides to form highly reactive metal nitrene intermediates.49,50 Mechanistic Studies. The stoichiometric reactions of Fischer carbene complexes of tungsten or chromium with azo deriva-

N

31

Figure 6. Structure of diaziridine 41 in the solid state

We conjectured that the new catalytic metathesis reaction of azobenzene derivatives described herein might commence by attack of an N-atom of the −N=N− bridge onto the half-

ACS Paragon Plus Environment

Page 7 of 10 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of the American Chemical Society

sandwich rhodium (iridium) carbene to form an ylide intermediate in the first place. With samples of pure crystalline carbenes in hand, we were in a positon to test this hypothesis. To this end, a sample of (Z)-16 was prepared according to the literature,51 as the preparative results had clearly indicated that only this geometric isomer engages in metathetic N=N bond cleavage. Slow warming of a 1:1 mixture of (Z)-16 and complex 10c in CD2Cl2 solution in the dark from −78°C to ambient temperature afforded the expected α-imino ester 31 and unreacted azobenzene in a ≈ 2.9:1 ratio (NMR) (Scheme 5); importantly, recovered 16 had completely isomerized to the thermodynamically more stable (E)-isomer. This outcome shows that the reaction pathway involves a reversible step generating the unreactive azobenzene isomer and hence explains why constant light-driven E→Z isomerization is necessary for full conversion in a catalytic set-up. Treatment of (Z)16 with two equivalents of 10c led to clean formation of 31 as the only product detected by NMR. This experiment proved that both halves of the azobenzene convert into product; the use of the unsymmetrical substrate 38 confirmed this conclusion (Scheme 4). Monitoring of the 1:1 mixture of 10c and 16 by NMR showed the build-up of an intermediate, which reached maximum concentration at −40°C; it proved stable enough for full characterization when kept cold. This compound was unambiguously assigned the diaziridine structure 41 based on extensive 1D and 2D NMR analyses; the trans-disposition of the two phenyl rings derived from the azobenzene was evident from the NOESY data and the presence of two signals in the 15N NMR spectrum. This stereochemical setting implies that cis/trans-isomerization must have taken place prior to formation of the three-membered ring; the cyclization is therefore almost certainly a stepwise rather than concerted process. The assignment was confirmed when crystals of this metastable intermediate suitable for X-ray diffraction could be grown. The structure of 41 in the solid state (Figure 6) features an almost isosceles trianglar core, in which the N1−N2 single bond (1.481(2) Å) is only slightly longer than C1−N1 (1.459(2) Å) and C1−N2 (1.463(3) Å) bonds and all three angles are very close to 60°. Diaziridine 41 reacts with a second equivalent of the rhodium half-sandwich carbene complex 10c to yield two equivalents of 3 (Scheme 5). This result suggests that the N-atoms of 41 are sufficiently nucleophilic to attack the carbene center. Diaziridines had been formed in modest yields in some of the stoichiometric reactions of Fischer tungsten carbenes with (electron deficient) azo compounds, when carried out in polar solvents such as MeCN.37a,52 Although a possible role in the stoichiometric metathesis of azobenzene had not been excluded, the −N=N− cleavage was explained in a canonical fashion via metal-nitrene complexes and diazametallacyclobutane derivatives in analogy to the Chauvin cycle of regular olefin metathesis.37a Actually, it was insinuated at one point that diaziridine formation and metathesis might be competing reaction channels.52 In striking contrast to the literature, all available experimental data suggest that the rhodium catalyzed cleavage of azobenzenes described herein passes through a diaziridine as the key intermediate. We confidently propose that the reaction proceeds as shown in Scheme 6: the electrophilic rhodium carbene B primarily formed succumbs to attack by an N-atom of

azobenzene, provided the latter is (Z)-configured. This step is likely reversible even at low temperature and leads to isomerization of the substrate, which therefore constantly needs to be photo-isomerized back to the active form to ensure complete conversion in the catalytic set-up. The resulting ylide F evolves into diaziridine G, such that the transannular interactions between the aryl substituents are minimized. Attack by a second carbene B affords an intermediate of type H which breaks down, presumably via a cheletropic process, to give two equivalents of the α-imino ester product.53 Overall, the stepwise scission of the −N=N− bond of azobenzene by a metal carbene described herein is reminiscent of the equally stepwise metathetic cleavage of the −N≡N triple bond of aryldiazonium salts by metal alkylidynes recently described by our group.54 Though certainly overshadowed by the immense success of alkene55,56 and alkyne metathesis,57 these and other unorthodox manifestations showcase that metathesis in general continues to provide many opportunities for methodological innovation.58 Scheme 6. Proposed Catalytic Cycle; [Rh] = Cp*RhI2

CONCLUSIONS Although the net outcome of the reaction between a rhodium carbene and the hetero-double bond of azobenzene formally represents a “metathesis” reaction, the underlying mechanism is distinct from the principles governing alkene metathesis, which became immensely popular and versatile with the advent of the well-defined Schrock- and Grubbs-type catalysts.55 In any case, the new manifold increases the portfolio of multiple-bonded substrates amenable to metathesis and provides access to synthetically valuable α-imino esters, some of which would be difficult to make otherwise. The most effective and tolerant precatalysts are half-sandwich complexes of type [Cp*MX2]2 (M = Rh, Ir), which perform better than the popular dirhodium tetracarboxylate complexes that dominate contemporary rhodium carbene chemistry. A representative set of pertinent half-sandwich Cp*M carbene complexes has been isolated and fully characterized, despite

ACS Paragon Plus Environment

Journal of the American Chemical Society 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

their sensitivity. The acquired data explain why the reactivity of this previously unexplored class of intermediates is strongly correlated with the nature of the halide ligands X; most notably, the complex derived from [Cp*RhCl2]2 undergoes spontaneous migratory insertion of the incipient carbene unit into the Rh−Cl bond before true carbene reactivity can be harnessed. This and related results allow a host of preparative data to be rationalized, and should provide guidance for further explorations of the field.

ASSOCIATED CONTENT Supporting Information. Experimental part including characterization data, NMR spectra of new compounds, and supporting crystallographic information (PDF)

AUTHOR INFORMATION Corresponding Author * [email protected]

ACKNOWLEDGMENT We thank Mr. S. Größl and Dr. J. Murphy for valuable discussions and the analytical departments of our Institute for excellent support. Generous financial support by the MPG is gratefully acknowledged.

REFERENCES (1) Doyle, M. P.; McKervey, M. A.; Ye, T. Modern Catalytic Methods for Organic Synthesis with Diazo Compounds: From Cyclopropanes to Ylides; Wiley: New York, 1998 (2) Modern Rhodium Catalyzed Organic Reactions; Evans, P. A., Ed.; Wiley-VCH: Weinheim, 2005. (3) (a) Davies, H. M. L.; Beckwith, R. E. J. Chem. Rev. 2003, 103, 2861. (b) Davies, H. M. L.; Manning, J. R. Nature 2008, 451, 417. (c) Davies, H. M. L. ; Morton, D. Chem. Soc. Rev. 2011, 40, 1857. (d) Davies, H. M. L.; Lian, Y. Acc. Chem. Res. 2012, 45, 923. (e) Davies, H. M. L.; Antoulinakis, E. G. Org. React. 2001, 57, 1. (4) (a) Doyle, M. P. Chem. Rev. 1986, 86, 919. (b) Doyle, M. P.; Forbes, D. C. Chem. Rev. 1998, 98, 911. (c) Doyle, M. P.; Duffy, R.; Ratnikov, M.; Zhou, L. Chem. Rev. 2010, 110, 704. (5) (a) Padwa, A.; Weingarten, M. D. Chem. Rev. 1996, 96, 223. (b) Padwa, A. Chem. Soc. Rev. 2009, 38, 3072. (6) (a) Ye, T.; McKervey, M. A. Chem. Rev. 1994, 94, 1091. (b) Taber, D. F.; Stiriba, S.-E. Chem. Eur. J. 1998, 4, 990. (c) Lebel, H.; Marcoux, J.-F.; Molinaro, C.; Charette, A. B. Chem. Rev. 2003, 103, 977. (d) Gois, P. M. P.; Afonso, C. A. M. Eur. J. Org. Chem. 2004, 3773. (e) Zhang, Z.; Wang, J. Tetrahedron 2008, 64, 6577. (f) Zhang, Y.; Wang, J. Coord. Chem. Rev. 2010, 254, 941. (g) Zhu, S.-F.; Zhou, Q.-L. Acc. Chem. Res. 2012, 45, 1365. (h) Gillingham, D.; Fei, N. Chem. Soc. Rev. 2013, 42, 4918. (i) Guo, X.; Hu, W. Acc. Chem. Res. 2013, 46, 2427. (j) Murphy, G. K.; Stewart, C.; West, F. G. Tetrahedron 2013, 69, 2667. (k) Ford, A.; Miel, H.; Ring, A.; Slattery, C. N.; Maguire, A. R.; McKervey, M. A. Chem. Rev. 2015, 115, 9981. (l) DeAngelis, A.; Panish, R.; Fox, J. M. Acc. Chem. Res. 2016, 49, 115. (7) For the preparation of metal carbenes from substrates other than diazo derivatives, see: (a) Jia, M.; Ma, S. Angew. Chem. Int. Ed. 2016, 55, 9134. (b) Gulevich, A. V.; Gevorgyan, V. Angew. Chem., Int. Ed. 2013, 52, 1371. (b) Davies, H. M. L.; Alford, J. S. Chem. Soc. Rev. 2014, 43, 5151. (c) Anbarasan, P.; Yadagiri, D. Rajasekar, S. Synthesis 2014, 46, 3004. (d) Wang, Y.; Lei, X.; Tang, Y. Synlett 2015, 26, 2051. (e) Barluenga, J.; Valdés, C. Angew. Chem., Int. Ed. 2011, 50, 7486. (f) Shao, Z.; Zhang, H. Chem. Soc. Rev. 2012, 41, 560. (8) For the pioneering work, see: Paulissen, R.; Reimlinger, H.; Hayez,, E.; Hubert, A. J.; Teyssié, P. Tetrahedron Lett. 1973, 14,

Page 8 of 10

2233. (b) Hubert, A. J.; Noels, A. F.; Anciaux, A. J.; Teyssié, P. Synthesis 1976, 600. (c) Demonceau, A.; Noels, A. F.; Hubert, A. J.; Teyssié, P. J. Chem. Soc. Chem. Commun. 1981, 688. (9) For leading studies on Rh(+3) carbenes, see the following and literature cited therein: (a) Callot, H. J.; Piechocki, Tetrahedron Lett. 1980, 21, 3489. (b) Callot, H. J.; Metz, F.; Piechocki, C. Tetrahedron 1982, 38, 2365. (c) O’Malley, S.; Kodadek, T. Tetrahedron Lett. 1991, 32, 2445. (d) Hayashi, T.; Kato, T.; Kaneko, T.; Asai, T.; Ogoshi, H. J. Organomet. Chem. 1994, 473, 323. (e) Lo, V. K.-Y.; Thu, H.-Y.; Chan, Y.-M.; Lam, T.-L.; Yu, W.-Y. ; Che, C.-M. Synlett 2012, 23, 2753. (f) Franssen, N. M. G.; Finger, M.; Reek, J. N. H.; de Bruin, B. Dalton Trans. 2013, 42, 4139. (10) Chen, W.-W.; Lo, S.-F.; Zhou, Z.; Yu, W.-Y. J. Am. Chem. Soc. 2012, 134, 13565. (11) (a) Shi, Z.; Koester, D. C.; Boultadakis-Arapinis, M.; Glorius, F. J. Am. Chem. Soc. 2013, 135, 12204. (b) Hyster, T. K.; Ruhl, K. E.; Rovis, T. J. Am. Chem. Soc. 2013, 135, 5364. (c) Liang, Y.; Yu, K.; Li, B.; Xu, S.; Song, H.; Wang, B. Chem. Commun. 2014, 50, 6130. (d) Jeong, J.; Patel, P.; Hwang, H.; Chang, S. Org. Lett. 2014, 16, 4598. (e) Hu, F.; Xia, Y.; Ye, F.; Liu, Z.; Ma, C.; Zhang, Y.; Wang, J. Angew. Chem., Int. Ed. 2014, 53, 1364. (f) Ye, B.; Cramer, N. Angew. Chem., Int. Ed. 2014, 53, 7896. (g) Shi, J.; Yan, Y.; Li, Q.; Xu, H. E.; Yi, W. Chem. Commun. 2014, 50, 6483. (h) Yang, Y.; Wang, X.; Li, Y.; Zhou, B. Angew. Chem., Int. Ed. 2015, 54, 15400. (i) Zhou, J.; Shi, J.; Liu, X.; Jia, J.; Xu, H. E.; Yi, W. Chem. Commun. 2015, 51, 5868. (j) Iagafarova, I. E.; Vorobyeva, D. V.; Peregudov, A. S.; Osipov, S. N. Eur. J. Org. Chem. 2015, 4950. (k) Wang, L.; Li, Z.; Qu, X.; Peng, W.-M.; Hu, S.-Q.; Wang, H.-B. Tetrahedron Lett. 2015, 56, 6214. (l) Ai, W.; Yang, X.; Wu, Y.; Wang, X.; Li, Y.; Yang, Y.; Zhou, B. Chem.−Eur. J. 2014, 20, 17653. (m) Zhang, Y.; Zheng, J.; Cui, S. J. Org. Chem. 2014, 79, 6490. (n) Yu, S.; Liu, S.; Lan, Y.; Wan, B.; Li, X. J. Am. Chem. Soc. 2015, 137, 1623. (o) Cheng, Y.; Bolm, C. Angew. Chem., Int. Ed. 2015, 54, 12349. (p) Zhou, B.; Chen, Z.; Yang, Y.; Ai, W.; Tang, H.; Wu, Y.; Zhu, W.; Li, Y. Angew. Chem., Int. Ed. 2015, 54, 12121. (q) Lu, Y.-S.; Yu, W.-Y. Org. Lett. 2016, 18, 1350. (r) Zhang, B.; Li, B.; Zhang, X.; Fan, X. Org. Lett. 2017, 19, 2294. (s) Halskov, K. S.; Roth, H. S.; Ellman, J. A. Angew. Chem. 2017, 129, 9311. (12) (a) 2.5:1 mixture of two isomers 3/3’, see the SI; the major isomer crystallized from the solution; (b) for a complex similar to 4, see: Qi, Z.; Yu, S.; Li, X. Org. Lett. 2016, 18, 700. (13) For applications that likely involve such species, see: (a) Qiu, L.; Huang, D.; Xu, G.; Dai, Z.; Sun, J. Org. Lett. 2015, 17, 1810. (b) Ng, F.-N.; Lau, Y.-F.; Zhou, Z.; Yu, W.-Y. Org. Lett. 2015, 17, 1676. (14) (a) Seidel, G.; Fürstner, A. Angew. Chem., Int. Ed. 2014, 53, 4807. (b) Seidel, G.; Gabor, B.; Goddard, R.; Heggen, B.; Thiel, W.; Fürstner, A. Angew. Chem., Int. Ed. 2014, 53, 879. (c) Seidel, G.; Mynott, R.; Fürstner, A. Angew. Chem., Int. Ed. 2009, 48, 2510. (d) Debrouwer, W.; Fürstner, A. Chem. Eur. J. 2017, 23, 4271. (e) Fürstner, A.; Davies, P. W. Angew. Chem. Int. Ed. 2007, 46, 3410. (f) Fürstner, A. Angew. Chem., Int. Ed. 2014, 53, 8587. (15) (a) Werlé, C.; Goddard, R.; Fürstner, A. Angew. Chem., Int. Ed. 2015, 54, 15452. (b) Werlé, C.; Goddard, R.; Philipps, P.; Farès, C. Angew. Chem. Int. Ed. 2016, 55, 10760. (16) (a) Leutzsch, M.; Wolf, L. M.; Gupta, P.; Fuchs, M.; Thiel, W.; Farès, C.; Fürstner, A. Angew. Chem. Int. Ed. 2015, 54, 12431. (b) D.-A. Roşca, K. Radkowski, L. M. Wolf, M. Wagh, R. Goddard, W. Thiel, A. Fürstner, J. Am. Chem. Soc. 2017, 139, 2443. (17) Werlé, C.; Goddard, R.; Philipps, P.; Farès, C.; Fürstner, A. J. Am. Chem. Soc. 2016, 138, 3797. (18) Review on iridium carbene chemistry, see: (a) Schafer, A. G.; Blakey, S. B. Chem. Soc. Rev. 2015, 44, 5969. For a recent advance, see: (b) Weldy, N. M.; Schafer, A. G.; Owens, C. P.; Herting, C. J.; Varela-Alvarez, A.; Chen, S.; Niemeyer, Z.; Musaev, D.G.; Sigman, M. S.; Davies, H. M. L.; Blakey, S. B. Chem. Sci. 2016, 7, 3142. (19) (a) Russel, A. E.; Brekan, J.; Gronenberg, L.; Doyle, M. P. J. Org. Chem. 2004, 69, 5269. (b) Doyle, M. P.; Hu, W.; Timmons, D. J. Org. Lett. 2001, 3, 933. (c) Davies, H. M. L.; DeMeese, J. Tetrahedron Lett. 2001, 42, 6803. For an interesting extension, see: (d) Padin, D.; Cambeiro, F.; Fananas-Mastral, M.; Varela, J. A.; Saá, C. ACS Catal. 2017, 7, 992.

ACS Paragon Plus Environment

Page 9 of 10 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of the American Chemical Society

(20) (a) Wang, Z.; Wen, J.; Bi, Q.-W.; Xu, X.-Q.; Shen, Z.-Q.; Li, X.-X.; Chen, Z. Tetrahedron Lett. 2014, 55, 2969. (b) Chai, G.-L.; Han, J.-W.; Wong, H. N. C. Synthesis 2017, 49, 181. (21) Liu, Y. Curr. Org. Chem. 2016, 20, 19. (22) For a more involved process in which transient rhodium carbenes react with (chiral) thioethers to give ylides that then engage in (asymmetric) epoxide formation, see: Aggarval, V. K:, Winn, C. L. Acc. Chem. Res. 2004, 37, 611. (23) In line with literature data,19,20 aliphatic aldehydes and acetophenone proved unsuitable under the chosen conditions. (24) Kornecki, K. P.; Briones, J. F.; Boyarskikh, Y.; Fullilove, F.; Autschbach, J.; Schrote, K. E.; Lancaster, K. M.; Davies, H. M. L.; Berry, J. F. Science 2013, 342, 351. (25) Nakamura, E.; Yoshikai, N.; Yamanaka, M. J. Am. Chem. Soc. 2002, 124, 7181. (26) Berry, J. F. Dalton Trans. 2012, 41, 700. (27) A related chloride migration to a carbene center is known in the platinum series, see: Bergamini, P.; Costa, E.; Sostero, S.; Orpen, A. G.; Pringle, P. G. Organometallics 1991, 10, 2989. (28) Attack of iodide onto a transient Rh(+3) acceptor carbene bearing a porphyrine ligand had been proposed to lead to a related complex, the structure of which was tentatively assigned based on NMR; this species failed to react with styrene, see: Maxwell, J.; Kodadek, T. Organometallics 1991, 10, 4. (29) For comparison, see a C-metalated Rh(+3) enolate prepared by a different route: Wu, J.; Bergman, R. G. J. Am. Chem. Soc. 1989, 111, 7628. (30) For recent advances in the use of donor/donor rhodium carbenes, see: (a) Soldi, C.; Lamb, K. N.; Squitieri, R. A.; GonzálezLópez, M.; Di Maso, M. J.; Shaw, J. T. J. Am. Chem. Soc. 2014, 136, 15142. (b) Lamb, K. N.; Squitieri, R. A:; Chintala, S. R.; Kwong, A. J.; Balmond, E. I.; Soldi, C.; Dmitrenko, O.; Castineira Reis, M.; Chung, R.; Addison, J. B.; Fettinger, J. C.; Hein, J. E.; Tantillo, D. J.; Fox, J. M.; Shaw, J. T. Chem. Eur. J. 2017, 23, 11843. (31) For remotely related carbene complexes of Cp*M(+3) stabilized by a tethered phosphine, see: Campos, J.; Carmona, E. Organometallics 2015, 34, 2212. (32) For an [(cod)IrCl]2 catalyzed insertion of acceptor/acceptor carbenes into water or N−H bonds, see: Ramakrishna, K.; Sivasankaar, C. Org. Biomol. Chem. 2017, 15, 2392. (33) For applications of carbene complexes derived from [(cod)IrCl]2, see the following and literature cited therein: (a) Lebel, H.; Ladjel, C. Organometallics 2008, 27, 2676. (b) Vaitla, J.; Beyer, A.; Hopmann, K. H. Angew. Chem. Int. Ed. 2017, 56, 4277. (34) Fairly strong back-donation was noticed in half-sandwich complexes of CpM(+1) (M = Rh, Ir) with a donor/donor carbene backbone, see: (a) Werner, H.; Schwab, P.; Bleuel, E.; Mahr, N.; Steinert, P.; Wolf, J. Chem. Eur. J. 1997, 3, 1375. (b) Werner, H.; Schwab, P-: Bleuel, E.; Mahr, N.; Windmüller, B.; Wolf, J. Chem. Eur. J. 2000, 6, 4461. (c) Ortmann, D. A.; Weberndörfer, B.; Ilg, K.; Laubender, M.; Werner, H. Organometallics 2002, 21, 2369. (35) (a) For a half-sandwich acceptor/acceptor complex of Ir(+1), see: Yuan, J.; Hughes, R. P.; Rheingold, A. L. Eur. J. Inorg. Chem. 2007, 4723. (b) dichlorocarbene complexes of Ir(+3), see: Clark, G. R.; Roper, W. R.; Wright, A. H. J. Organomet. Chem. 1982, 236, C7. (36) (a) Hegedus, L. S.; Kramer, A. Organometallics 1984, 3, 1263. (b) Hegedus, L. S.; Lundmark, B. R. J. Am. Chem. Soc. 1989, 111, 9194. (37) (a) Tata Maxey, C.; Sleiman, H. f.; Massey, S. T.; McElweeWhite, L. J. Am. Chem. Soc. 1992, 114, 5153. (b) Sleiman, H. F.; McElwee-White, L. J. Am. Chem. Soc. 1988, 110, 8700. (c) Sleiman, H. F.; Mercer, S.; McElwee-White, L. J. Am. Chem. Soc. 1989, 111, 8007. (d) Massey, S. T.; Barnett, N. D. R.; Abboud, K. A.; McElweeWhite, L. Organometallics 1996, 15, 4625. (38) For the related metathesis of nitroso compounds, see: (a) Pilato, R. S.; Williams, G. D.; Geoffroy, G. L.; Rheingold, A. L. Inorg. Chem. 1988, 27, 3665. (b) Herndon, J. W.; McMullen, L. A. J. Organomet. Chem. 1989, 368, 83. (c) Peng, W.-J.; Gamble, A.S.; Templeton, J. L.; Brookhart, M. Inorg. Chem. 1990, 29, 463. (39) Eftekhari-Sis, B.; Zirak, M. Chem. Rev. 2017, 117, 8326.

(40) Dickstein, J. S.; Fennie, M. W.; Norman, A. L.; Paulose, B. J.; Kozlowski, M. C. J. Am. Chem. Soc. 2008, 130, 15794. (41) For a leading reference, see: Palacios, F.; Vicario, J.; Aparicio, D. J. Org. Chem. 2006, 71, 7690. (42) Mandler, M. D.; Truong, P. M.; Zavaliij, P. Y.; Doyle, M. P. Org. Lett. 2014, 16, 740. (43) (a) Gu, P.; Wu, X.-P.; Su, Y.; Li, X.-Q.; Xue, P.; Li, R. Synlett 2014, 535. (b) See also: Huang, H.; Wang, Y.; Chen, Z.; Hu, W. H. Synlett 2005, 2498. (c) Qian, Y.; Jing, C.; Zhai, C.; Hu, W.-H. Adv. Synth. Catal. 2012, 354, 301. (44) For unrelated and mechanistically obscure cleavage of azobenzene by stoichiometric rhodium, see: Bruce, M. I.; Goodall, B. L.; Iqbal, M. Z.; Stone, F. G. A. J. Chem. Soc. D 1971, 661. (45) (a) Sharma, S.; Han, S.-H.; Han, S.; Ji, W.; Oh, J.; Lee, S.-Y.; Oh, J. S.; Jung, Y. H.; Kim, I. S. Org. Lett. 2015, 17, 2852. (b) Son, J.-Y.; Kim, S.; Jeon, W. H.; Lee, P. H. Org. Lett. 2015, 17, 2518. (c) see also: Long, Z.; Wang, Z.; Zhou, D.; Wan, D.; You, J. Org. Lett. 2017, 19, 2777. (46) In the photostationary state, the isomer ratio of azobenzene is Z:E = 37:63, cf.; Fischer, E.; Frankel, M.; Wolovsky, R. J. Chem. Phys. 1955, 23, 1367. (47) Bandara, H. M. D.; Burdette, S. C. Chem. Soc. Rev. 2012, 41, 1809. (48) Oro, L. A.; Carmona, D.; Lahoz, F. J.; Puebla, M. P.; Esteban, M.; Foces-Foces, C.; Cano, F. H. J. Chem. Soc., Dalton Trans. 1986, 2113. (49) See the following for leading references and literature cited therein: (a) Driver, T. G. Org. Biomol. Chem. 2010, 8, 3831.(b) Stokes, B. J.; Dong, H.; Leslie, B. E.; Pumphrey, A. L.; Driver, T. G. J. Am. Chem. Soc. 2007, 129, 7500. (c) Stokes, B. J.; Jovbanovic, B.; Dong, H.; Richert, K. J.; Riell, R. D.; Driver, T. G. J. Org. Chem. 2009, 74, 3225. (d) Stokes, B. J.; Richert, K. J.; Driver, T. G. J. Org. Chem. 2009, 74, 6442. (e) Shen, M.; Leslie, B. E.; Driver, T. G. Angew. Chem. Int. Ed. 2008, 47, 5056. (50) Even if nitrene formation were slower than carbene formation, the resulting dirhodium carbene species would instantly react with the azide, see ref 42 and the following: (a) Wee, A. G. H.; Slobodian, J. J. Org. Chem. 1996, 61, 2897. (b) Blond, A.; Moumné, R.; Bégis, G.; Pasco, M.; Lecourt, T.; Micouin, L. Tetrahedron Lett. 2011, 52, 3201. (51) Cook, A. H. J. Chem. Soc. 1938, 876. (52) Tata Maxey, C.; McElwee-White, L. Organometallics 1991, 10, 1913. (53) Diaziridines undergo ring opening when on reaction with electrophilic species such as ketenes, cf.: Komatsu, M.; Nishikaze, N.; Sakamoto, M.; Ohshiro, Y.; Agawa, T. J. Org. Chem. 1974, 39, 3198. (54) Lackner, A. D.; Fürstner, A. Angew. Chem. Int. Ed. 2015, 54, 12814. (55) (a) Schrock, R. R. Angew. Chem. Int. Ed. 2006, 45, 3748. (b) Grubbs, R. H. Angew. Chem. Int. Ed. 2006, 45, 3760. (c) Chauvin, Y. Angew. Chem. Int. Ed. 2006, 45, 3740. (56) (a) Fürstner, A. Angew. Chem. Int. Ed. 2000, 39, 3012. (b) Blechert, S.; Connon, S. J. Angew. Chem. Int. Ed. 2003, 42, 1900. (c) Nicolaou, K. C.; Bulger, P. G.; Sarlah, D. Angew. Chem. Int. Ed. 2005, 44, 4490; (d) Fürstner, A. Chem. Commun. 2011, 47, 6505. (e) Fürstner, A. Science 2013, 341, 1229713. (f) Hoveyda, A. H. J. Org. Chem. 2014, 79, 4763. (57) Fürstner, A. Angew. Chem. Int. Ed. 2013, 52, 2794. (58) For some unorthodox manifestations of “metathesis”, see: Nitriles: (a) Geyer, A. M.; Wiedner, E. S.; Gray, J. B.; Gdula, R. L.; Kuhlmann, N. C.; Johnson, M. J. A.; Dunietz, B. D.; Kampf, J. W. J. Am. Chem. Soc. 2008, 130, 8984. Imines: (b) Zuckerman, R. L.; Krska, S. W.; Bergman, R. G. J. Am. Chem. Soc. 2000, 122, 751. (c) Cantrell, G. K.; Meyer, T. Y. J. Am. Chem. Soc. 1998, 120, 8035. Alkanes: (d) Basset, J. M.; Copéret, C.; Lefort, L.; Maunders, B. M.; Maury, O.; Le Roux, E.; Saggio, G.; Soignier, S.; Soulivong, D.; Sunley, G. J.; Taoufik, M.; Thiovolle-Cazat, J. J. Am. Chem. Soc. 2005, 127, 8604. (e) Goldman, A. S.; Roy, A. H.; Huang, Z.; Ahuja, R.; Schinski, W.; Brookhart, M. Science 2006, 312, 257. Metathesis of C−S and C−P bonds: (f) Liang, Z.; Bhawal, B. N.; Yu, P.; Morandi, B. Science 2017, 356, 1059.

ACS Paragon Plus Environment

Journal of the American Chemical Society 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Paragon Plus Environment

Page 10 of 10